Sei sulla pagina 1di 15

Geophys. J. Int. (2010) 181, 275–289 doi: 10.1111/j.1365-246X.2010.04537.

GJI Geomagnetism, rock magnetism and palaeomagnetism


Rock magnetic evidence for inclination shallowing in the early
Carboniferous Deer Lake Group red beds of western Newfoundland

Dario Bilardello∗ and Kenneth P. Kodama


Department of Earth and Environmental Sciences, Lehigh University, 31 Williams Dr., Bethlehem, PA, USA. E-mail: bilardello@geophysik.uni-muenchen.de

Accepted 2010 January 20. Received 2010 January 19; in original form 2008 July 4

SUMMARY
A paleomagnetic and rock magnetic study of the Carboniferous Deer Lake Group red beds of
Newfoundland was performed to detect and correct for inclination shallowing. Results indicate
a primary remanence carried by magnetite, with a mean direction of D = 179.7◦ , I = 33.7◦ ,
α 95 = 7.2◦ which corresponds to a paleopole position of 22.2◦ N, 122.3◦ E, A95 = 7.6◦ . Correct-
ing the inclination using anisotropy of anhysteretic remanence and the measured individual
particle anisotropy gives a corrected direction of D = 178.8◦ , I = 50.9◦ , α 95 = 6.3◦ cor-
responding to a paleopole position at 8.4◦ N, 122.7◦ E, A95 = 7.2◦ . This correction is larger
than that of other red beds from the Maritime Provinces of Canada, but is consistent with
paleoenvironmental reconstructions, placing North America in a more arid climate zone. Our
inclination-corrected results have important implications for this portion of North America’s
apparent polar wander path and suggest a correction is needed for other red bed-derived
APWPs. We have determined the range of flattening factors f , defined as the proportional-
ity constant between the tangents of the measured (I m ) and field (I o ) inclinations, tan(I m ) =
f tan(I 0 ), from this study and previous inclination correction studies to estimate inclination cor-
rections. Using the range of haematite f factors observed in this study to correct the Neogene
red bed inclinations from the Vallès-Penedès Basin (NE Spain) yields inclinations consistent
with the known geomagnetic field inclination in the Neogene, thus indicating that the range
of f factors reported here may be used to estimate the magnitude of inclination shallowing in
red beds.
Key words: Magnetic fabrics and anisotropy; Magnetic mineralogy and petrology; Palaeo-
magnetism applied to tectonics; Palaeomagnetism applied to geologic processes; Rock and
mineral magnetism.

paleomagnetic study (Roy & Park 1972; Walker et al. 1981; Larson
1 I N T RO D U C T I O N
et al. 1982; Liebes & Shive 1982).
The stable remanence of red beds has long been studied by paleo- In some cases, the remanence is carried by detrital haematite,
magnetists because of red beds’ frequent occurrence in continental making the remanence a primary detrital remanent magnetization
sedimentary records. The apparent polar wander paths (APWPs) (DRM) (Collinson 1974; Elston & Purucker 1979; Steiner 1983;
of some continents, for instance, such as the Meso-Paleozoic por- Tauxe et al. 1990; Tan & Kodama 2002; Tan et al. 2007; Bilardello
tion of North America’s APWP, are dominated by red bed-derived 2008; Bilardello & Kodama 2009). In many red beds, however,
paleopoles (Van der Voo 1990; McElhinny & Lock 1993, 1996). different magnetic mineralogies often coexist and different combi-
The mechanism by which red beds acquire their remanence, and nations of primary and secondary mineralogies can occur, making
therefore the accuracy with which red beds record paleomagnetic di- paleomagnetic studies of red beds case specific. For instance, pig-
rections, is poorly understood (Collinson 1965; Dunlop & Özdemir mentary haematite is typically present in red beds, giving them
1997). The most common view is that the remanence is carried by their characteristic red colouration, and is secondary in origin
haematite of chemical origin, in which case the remanence will be a (Dunlop & Özdemir 1997; McElhinny & McFadden 1999; Kodama
chemical remanent magnetization (CRM) that was acquired either & Dekkers 2004). Other red beds may have a primary remanence
soon after or up to millions of years after deposition (Collinson carried by magnetite or maghemite, with pigmentary haematite (and
1974; Dunlop & Özdemir 1997). The timing of the acquisition of other mineralogies) present as secondary phases (see Dunlop &
the remanence is a fundamental parameter in the interpretation of a Özdemir 1997). For any of the earlier cases, the primary remanence
may be affected by depositional and/or post-depositional processes.

Now at: Ludwig Maximilians University, Theresienstrasse, 41, 80333 Burial compaction, which can shallow the inclination of the mag-
Munich, Germany. netization, is the focus of this paper (Anson & Kodama 1987; Sun


C 2010 The Authors 275
Journal compilation 
C 2010 RAS
276 D. Bilardello and K. P. Kodama

& Kodama 1992). Shallowing of the paleomagnetic inclination is an over 2000-m-thick fluvial and alluvial fan deposit (Belt 1969;
well-documented in red beds (Garcés et al. 1996; Gilder et al. 2001; Hyde 1979), composed of mainly red to grey arkosic sandstone and
Tan & Kodama 2002, 2003; Tauxe & Kent 2004; Kent & Tauxe conglomerate, with minor red to grey siltstone and pink to grey mi-
2005; Vaughn et al. 2005; Tan et al. 2007). Inclinations that are critic limestone. The Rocky Brook Formation lies conformably on
too shallow result in erroneous paleolatitude estimates and strongly top of the North Brook, and its lowermost layers intertongue with
affect the APWPs of continents, hence the need to correct for this facies equivalent to the North Brook Formation. It is typically 500-
error (Garcés et al. 1996; Gilder et al. 2001; Tan & Kodama 2003; to 600-m thick and consists of red calcareous silts, grey to green
Vaughn et al. 2005; Tan et al. 2007). siltstones and mudstones, calcareous dolostones, dolomitic lime-
Two inclination shallowing correction techniques have been pro- stones, and grey, green and black mudstones. The unit is interpreted
posed. One is based on magnetic anisotropy measurements (Jackson by Hyde (1984) to be mixed lacustrine and fluvial deposits. The
et al. 1991; Kodama 1997; Tan & Kodama 2002, 2003; Vaughn Humber Falls Formation rests disconformably on top of the Rocky
et al. 2005; Tan et al. 2007), the other is based on models of the Brook Formation and has a maximum thickness of 250 m. It has
geomagnetic field secular variation [the elongation–inclination (EI) been interpreted by Belt (1969) and Hyde (1979) to be an alluvial
technique] (Tauxe & Kent 2004; Kent & Tauxe 2005). Anisotropy- fan-fluvial deposit composed of grey to reddish conglomerate and
based corrections rely on the assumption that the orientation distri- arkosic sandstones and red to grey siltstones and mudstones. The
bution of magnetic particle easy axes in sedimentary rocks is nearly Little Pond Brook Formation reaches a thickness of 750 m and is
isotropic soon after deposition and acquires a bedding parallel foli- considered to be the fluvial (Hyde & Ware 1981) lateral equivalent
ated anisotropy during burial compaction (Stephenson et al. 1986; of the Humber Falls Formation. It is composed of grey to red clas-
Jackson et al. 1991; Tan & Kodama 2003). The magnetic fabric, tics and grey coarse-grained sandstone and pebble conglomerate
together with the individual particle anisotropy is used to correct interbedded with red fine-grained sandstone and siltstone. Spore
the inclination (Jackson et al. 1991; Tan & Kodama 2002, 2003; assemblages in the Deer Lake Group suggest a Visean to possibly
Tan et al. 2003). Corrections that are based on models of the geo- early Namurian age (Hyde 1984).
magnetic field, on the other hand, rely on the expected distribution Two northeast-trending synclines separated by an intervening
of paleomagnetic directions caused by secular variation. During a anticline deform the Deer Lake Group, the Humber syncline in the
period long enough to observe secular variation (∼103 years), pale- west and the Grand Lake syncline in the east. Bedding attitudes vary
omagnetic directions will be elongated in a N–S direction, with the from sub-horizontal to overturned. It has been suggested by Hyde
amount of elongation depending on the sampling site’s initial pale- (1979) that folding of the Deer Lake Group was the result of dip-slip
olatitude, with the maximum elongation occurring at the equator. movement along major northeasterly trending faults. These faults
Shallowed inclinations, however, will result in flattened distribu- controlled the overall form and sedimentation of the basin and may
tions that are elongate in an E–W direction. Tauxe & Kent (2004) have been active, with strike-slip movement along them, during the
have developed a correction technique that relies on models of the deposition and deformation of the older Anguille Group.
geomagnetic field to restore the elongation and inclination of the
distribution.
In this paper, we perform a paleomagnetic and rock magnetic
3 SAMPLING AND METHODS
study of the early Carboniferous Deer Lake Group red beds from
Newfoundland to detect and correct for inclination shallowing. Re- As Irving & Strong (1984) have pointed out, access is difficult and
sults are compared to other Carboniferous red bed-derived pale- exposures are few and often weathered. Sampling was performed
opoles from the Maritime Provinces of Canada that have been cor- using the collection sites of Irving & Strong (1984) as guidance,
rected for inclination shallowing (Bilardello 2008; Bilardello & however many sites are along stream beds and the water level was
Kodama 2010). The effects of the corrections on the Carbonifer- still too high for sampling during both field seasons of June and
ous APWP for North America are evaluated in terms of the pale- August of 2004. Nevertheless, 18 sites were sampled from all four
oenvironmental reconstructions of the rocks sampled. Finally, we units of the Deer Lake Group, mostly along lakeshores over a strati-
determine the relationship between corrected inclinations and the graphic thickness of about 2000 m (Fig. 1). Many sites were not
amount of shallowing for all available inclination-corrected results. ideal: they were either too weathered or coarse-grained, which made
We check the validity of this relationship by applying a simplified drilling very difficult and resulted in many cores breaking. As a re-
correction to the Neogene Vallès-Penedès Basin (NE Spain) red sult, many short cores were collected, so sampling sites ranged from
beds. ∼8 to ∼15 cores that yielded one to three samples each. Cores were
collected with a gasoline-powered 25-mm-diameter diamond coring
drill and oriented using a magnetic compass.
The magnetic mineralogy of the Deer Lake group rocks was in-
2 GEOLOGY
vestigated by a variety of techniques. IRM acquisition experiments
The Early Carboniferous Deer Lake Group is part of the fresh water in fields up to 5 T applied with an ASC Scientific impulse magne-
sedimentary rocks that fill the Deer Lake Basin and rest on Pre- tizer were used to identify the major coercivity components. IRM
cambrian (Grenville Structural Province of the Canadian Shield) acquisition modelling was conducted (Kruiver et al. 2001). Thermal
basement gneisses. The deposits of the Deer Lake Basin have been demagnetization of composite IRMs of 0.1, 0.3 and 1.2 T was per-
subdivided by Hyde (1984) into the Tournaisian Anguille Group, formed using an ASC TD-48 thermal demagnetizer. These exper-
the Wigwam Brook and Wetstone Point Formations, the uncon- iments were used to identify magnetic mineralogy (Lowrie 1990).
formably overlying predominantly Visean Deer Lake Group and Magnetic remanence was measured with a 2G Enterprises supercon-
the Westphalian Howley Group (Fig. 1). ducting magnetometer at Lehigh University. Low-temperature mag-
The four units of the Deer Lake Group are well exposed and are netic properties were measured with a Quantum Designs MPMS
best suited for a paleomagnetic study. All four units were sampled cryogenic susceptometer at the Institute for Rock Magnetism, Uni-
for this study. The lowermost unit, the North Brook Formation is versity of Minnesota. Specimens were cooled down to 20 K in zero


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 277

Figure 1. Location map of the Deer Lake Group of Newfoundland. Sampling sites are shown as circles on the map modified after Irving & Strong (1984). For
a colour version of the figure, the reader is referred to the online version of this paper.

field and the remanence was measured on heating to 300 K and demagnetization steps. The nine pARMs were used to determine
on cooling back to 20 K. First-order reversal curves (FORCs) were the best-fit anisotropy tensor by least squares. Finally, the individ-
measured with a variable temperature Princeton Measurements Cor- ual particle anisotropy, or a factor, was measured directly following
poration vibrating sample magnetometer at the Institute for Rock several steps. First, the samples were crushed, ball milled and fi-
Magnetism and processed using Winklhofer & Zimanyi’s (2006) nally sieved to ensure that the rock particles are smaller than 4 
MATLAB code. (62.5 μm). Then, a slurry was made with the resulting crushed rock
Thermal demagnetization was conducted in the thermal demag- and distilled water which was then placed in an ultrasonic cleaner for
netizer using 16 steps between 200 and 685 ◦ C to identify the at least 24 hr before circulating it in a magnetic extraction apparatus
characteristic remanence of the rocks. Chemical demagnetization (Hounslow & Maher 1999). The magnetic separate was collected
was conducted by leaching the samples in 3N HCl using 4-day over a period of 2–3 weeks. Small amounts of the magnetic separates
leaching steps up to a cumulative 12 days and then 7-day steps were mixed in a slow-drying epoxy resin for 24 hr and placed in a
up to a cumulative 40 days. AF demagnetization was conducted dc magnetic field (45 and 55 mT) to align the grains. Because the
on representative samples in steps of 5 mT up to 50 mT and then grains are aligned, the bulk AAR anisotropy of the epoxy samples
10 mT steps up to 100 mT. provides an average individual particle anisotropy.
Determining the remanence anisotropy of the characteristic ChRMs were corrected for inclination shallowing using the rela-
remanence-carrying grains involved several steps. First, the coerciv- tionship between magnetite anisotropy principal axes and inclina-
ity spectra of the ferromagnetic grains was determined from partial tion error described by Jackson et al. (1991):
anhysteretic remanence magnetization (pARM) using a Schonstedt
qx (a + 2) − 1
GSD-5 tumbling specimen demagnetizer modified to apply pARMs. tan(I0 ) = tan(Im ), (1)
pARM acquisition curves were acquired in 10 mT intervals between qz (a + 2) − 1
0 and 100 mT to determine the median coercivity window. Second, where I 0 is the corrected paleomagnetic inclination; I m is the mea-
the magnetic fabric of these grains was measured by anisotropy sured paleomagnetic inclination; qx and qz are the normalized max-
of anhysteretic remanence (AAR) in a 10–70 mT coercivity win- imum and minimum axes of the magnetic anisotropy ellipsoid
dow, with a 0.1 mT dc field. AAR was determined following the (qx + qy + qz = 1); a is the individual particle anisotropy (the
method described by McCabe et al. (1985): pARMs were applied a factor = easy magnetic axis/hard magnetic axis). An inclination
in nine different orientations separated by 100 mT alternating field correction may use either AMS or AAR parameters (both anisotropy


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
278 D. Bilardello and K. P. Kodama

principal axes and particle anisotropy), because the magnitudes of clearly shows the 119 K transition, although not as abruptly. Both
the corresponding normalized principal axes are linearly correlated transitions are best observed when plotting the derivatives of the
(Stephenson et al. 1986; Tan & Kodama 2002, 2003; Tan et al. two curves (Fig. 4).
2003). Inclinations have been corrected using the measured a factor Hysteresis loops were acquired in fields of 1.2 T to capture a
and the measured anisotropy principal axes. broad spectrum of coercivities (Fig. 5a), however, the FORC dia-
gram of specimen NF11j-2, that has representative behaviour, shows
the magnetic mineralogy being dominated by one low (<100 mT)
4 R E S U LT S coercivity phase or group of phases (Fig. 5b).

4.1 Magnetic mineralogy


4.2 Paleomagnetism
Thermal demagnetization of orthogonal IRMs of 0.1, 0.3 and 1.2
T (Lowrie 1990) reveals the presence of at least three components Thermal demagnetization of the Deer Lake Group rocks revealed
of magnetization (Fig. 2). The 1.2 T demagnetization curve shows different components of magnetization that are completely removed
a drop around 680 ◦ C. Typically this drop is small as in samples by demagnetization up to 600–625 ◦ C. The lowest unblocking tem-
NF11e2 and NF15e3, but for some samples, as in sample NF8e2 it perature component, denoted the A component, is completely re-
is the most prominent feature. In some cases, as for example sample moved by 200 ◦ C, but it is not possible to fully resolve it with our
NF8e2, it also shows a distinct decrease of intensity between 350 demagnetization scheme (Fig. 6).
and 450 ◦ C. The intermediate (0.3 T) demagnetization curve has An intermediate unblocking temperature component of magne-
different behaviour in different samples: in some cases it shows a tization, denoted B, is sometimes observed between 200 and 300–
distinct drop at 580 ◦ C after which it is mostly removed, as in sample 400 ◦ C (Fig. 6d). Although this component is not fully resolved
NF11e2, but in others it persists up to 680 ◦ C, as in samples NF8e2 and only defined by three demagnetization steps, it is southerly
and NF15e3, a slight decrease around 400–450 ◦ C is also observed. directed with positive inclinations (reverse polarity). The highest
The 0.1 T curve consistently shows a decrease at 350 ◦ C and then unblocking temperature component of magnetization, component
drops sharply at 580 ◦ C. Upon reapplication of IRMs in the same C, is observed above 300–400 ◦ C and is completely removed by
fields, but after complete thermal demagnetization, the intensities 600–625 ◦ C. This component has a northerly direction and negative
of all IRMs show a decrease with respect to their initial intensity. inclination (normal polarity) (Figs 6b and d). A few samples, how-
IRM acquisition modelling of specimens NF15e-1 and NF8e-1 ever, revealed an antipodal direction of component C that trends
revealed three major coercivity components. The lowest component towards the origin starting at 200 ◦ C and is completely removed by
is acquired between 60 and 80 mT. An intermediate component 600–625 ◦ C. This component is always southerly directed and has
peaks at 500 mT, whereas the highest component peaks around a positive inclination (Figs 6a and e). Complete demagnetization
2000 mT (Fig. 3). up to 685 ◦ C was performed on all samples, however no stable di-
The low-temperature measurements of specimen NF11j-1 clearly rection has been isolated above 600–625 ◦ C (Figs 6c and f). The
show an initial loss of intensity when heating from 20 to 50 K and Deer Lake Group mean direction for this high unblocking temper-
then a second abrupt drop at around 119 K. The cooling curve only ature component of magnetization in stratigraphic coordinates is

Figure 2. Thermal demagnetization of three orthogonal IRMs of 1.2 T (diamonds), 0.3 T (squares) and 0.1 T (triangles) (Lowrie 1990) for specimens NF8e2,
NF11e2 and NF15e3. Filled symbols are re-applications of the IRMs at the end of the heating cycle (specimens NF8E2 and NF11E2). Loss of intensity on
re-application of the IRMs indicates the presence of maghemite. Losses of intensity in the 350–450 ◦ C range is indicative of maghemite whereas drops of
intensity at 580 and 680 ◦ C indicate magnetite and haematite, respectively.


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 279

Figure 3. Modelling of IRM acquisition curves (Kruiver et al. 2001) for samples NF15e-1 and NF8e-1. The three peaks are in the coercivity ranges of
magnetite (low), haematite (intermediate) and goethite (high).

Figure 4. Low-temperature magnetization measurements for specimen NF11j-1 during heating from 20 to 300 K and cooling from 300 to 20 K all in zero
field. The inset shows the derivatives of the two curves to highlight magnetization transitions. The initial loss of intensity at <50 K during heating is attributed
to goethite. The Verwey transition, characteristic of magnetite is clearly visible around 119 K.

D = 179.3◦ , I = 30.5◦ , α 95 = 19.7◦ , N = 8. Chemical demagneti- interval for the fold test is extremely large, ranging from −58 to
zation was not successful at isolating stable components of magne- 151 per cent untilting. However, when the formation mean direc-
tization (Fig. 6g), so these results will not be treated. tion is plotted with its confidence ellipse together with the formation
For a few selected samples, AF demagnetization up to 50–60 mean direction isolated by Irving & Strong (1984), the two direc-
mT isolated the same components of magnetization as thermal de- tions are statistically indistinguishable (Fig. 7b). The directions iso-
magnetization (Fig. 6h). AF demagnetization up to 100 mT did not lated by Irving & Strong (1984) (Table 1) proved to be pre-folding
isolate a stable magnetization at higher fields (Fig. 6i). When the directions of magnetization. We have combined our site mean di-
site mean directions from thermal demagnetization (Table 1) are rections with those of Irving & Strong (1984) and together, the site
plotted on a stereographic projection there is much scatter (Fig. means pass the Tauxe & Watson (1994) and the McFadden & Jones
7a). Fold tests performed on these directions gave ambiguous re- (1981) fold tests, although with a large 95 per cent confidence
sults (Tauxe & Watson 1994), because the 95 per cent confidence interval (Fig. 8). The mean formation direction in stratigraphic


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
280 D. Bilardello and K. P. Kodama

Figure 5. FORC diagram obtained for specimen NF11j-2 using the code of Winklhofer & Zimanyi (2006). This sample was chosen because samples from
this site show typical thermal demagnetization behaviour. (a) Raw data and (b) FORC diagram (smoothing factor = 5) shows one main coercivity distribution
that is <100 mT dominating the magnetization. For a colour version of the figure, the reader is referred to the online version of this article.

coordinates obtained by combining our data (N = 8) with that of


Irving & Strong (1984) (N = 21) is D = 179.7◦ , I = 33.7◦ , α 95 = 4.4 Individual particle anisotropy (a factor)
7.2◦ . This direction has been used for our inclination correction.
Because Blow & Hamilton (1978) found a dependence of inclination
shallowing on porosity, DRM research has focused on the attach-
ment of magnetic particles to clays during DRM acquisition and the
4.3 Magnetic Fabrics
effect of the attachment on inclination shallowing (e.g. Scherbakov
AAR was used to measure the magnetic fabrics. To determine the and Scherbakova 1983; Anson & Kodama 1987; Deamer and
coercivity range to apply the ARMs, we applied pARMs between Kodama, 1990; Jackson et al. 1991; Mitra & Tauxe 2009). Flocs are
0 and 100 mT in intervals of 10 mT. The highest activation of thought to be formed by magnetic particles embedded in clay miner-
magnetic grains, indicated by a peak in the pARM spectrum, oc- als, which are in turn arranged in a hierarchical, fractal nature (Mitra
curred between 10 and 70 mT. We therefore measured the AAR & Tauxe 2009 and references therein). However, the true structure
in this coercivity range. Because the ChRM is demagnetized over and, most importantly, the orientation of individual magnetic parti-
the same AF interval (15–50/60 mT), the AAR is applied to the cles within the flocs is unknown. This raises an important question
ChRM-carrying grains. The magnetic remanence fabric in strati- when considering particle anisotropy for inclination shallowing be-
graphic coordinates (Fig. 9a) has a sub-bedding parallel foliation cause it carries the implication that the magnetic anisotropy of the
with very steep minimum axes plunging to the NW. The maximum flocs could be more important than that of the particles embedded
and intermediate axes show good clustering at shallow inclina- in them. However, the misalignment of magnetic grain easy axes
tions. The maximum axes define a strong magnetic lineation that in a floc, as well as the misalignment of mean floc “easy axes” is
plunges gently to the SSW, whereas the intermediate axes plunge measured as the overall preferred alignment of magnetic particles
to the ESE. Orientation of the mean eigenvectors of the fabric by the bulk rock anisotropy and the individual magnetic particle
are better appreciated after re-sampling the data using the boot- anisotropy, as measured by the a factor, is still the basic “quantum”
strap technique of Constable & Tauxe (1990) (Fig. 9b). Plotted on of magnetic anisotropy in the rock.
a Flinn diagram (Flinn 1962) (Fig. 9c) and as histograms of the Two samples of magnetic separates have been prepared, mixed
eigenvalues (Fig. 9f), the fabric appears triaxial tending towards in epoxy, and dried in dc magnetic fields of 45 and 55 mT.
oblate. AAR of the sample dried in a 55 mT field had an anisotropy of
In geographic coordinates (Fig. 9d, e), the clustering of all princi- 204 per cent (Rees 1966) and an RMS error of 3.4 per cent.
pal axes is less well defined. The minimum axes plunge towards the The a factor measured from this fabric is 5.39. This high value
SW at intermediate angles, whereas the maximum and intermediate is inconsistent with previous measurements of magnetite a factors
axes are scattered along a great circle oriented NW–SE and gently (Kodama 1997; Tan & Kodama 1998; Kodama 2009) and may be
dipping to the NE. due to clumping of the magnetic particles in the epoxy sample.


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 281

Figure 6. Orthogonal vector endpoint diagrams for thermal demagnetization (Zijderveld 1967). The ChRM was typically isolated between 200 and 600–
625 ◦ C. (a) and (e) show reverse polarity (south and down) components, whereas (b) and (d) show normal polarity (north and up) ChRM components; (g)
shows the unstable behaviour of a chemically demagnetized sample, which is directly comparable to the behaviour observed in (d). The sample in (d) also
reveals a poorly constrained lower unblocking temperature component that is antipodal to the ChRM and can be observed on the equal area projection between
200 and 400 ◦ C, solid (open) circles are lower (upper) hemisphere projections. (h) Shows a result from AF demagnetization which is directly comparable to
(e). To indicate that complete demagnetization was performed, but did not isolate a stable remanence above 600–625 ◦ C or 50–60 mT, in (c) and (f) complete
demagnetization is shown up to 685 ◦ C of the same samples as in (b) and (e), where in (i) demagnetization up to 100 mT of the same sample as in (h) is shown.
For all diagrams, solid (open) symbols are projections onto the horizontal (vertical) plane.

Confirmation of this interpretation is the observation that the ARM magnetization effect (GRM) because this would result in remanent
directions measured often diverged from the directions of the field directions in an orthogonal plane to the ARMs imparted (Dunlop &
used to impart the ARM, indicating strong magnetic interactions Özdemir 1997), whereas the measured remanence for all the nine
between the grains. We doubt the divergence of the imparted ARMs measurement positions used to calculate the anisotropy are subpar-
and the measured remanence is to be attributed to a gyroremanent allel to the orientation of the dc setting field. In contrast, the sample


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
282 D. Bilardello and K. P. Kodama

Table 1. Site mean directions isolated through thermal demagnetization. Sites nf1–15 are from this study,
sites 06–85 are from Irving & Strong (1984) with the family of magnetization indicated, in situ declinations
and inclinations are calculated from the reported directions in stratigraphic coordinates and bedding attitude,
R and α 95 (sites 22, 24, 67, 70, 71, 75) are calculated from the reported N and k.
Deer Lake Group ChRMs thermal demagnetization
Site In situ D In situ I Strike Dip Strat D Strat I N R k α 95
nf1 160.7 5.9 347 30 158.6 1.9 5 4.9705 135.7 6.6
nf8 71.3 5.2 32 90 38.8 −39.1 5 4.8025 20.3 17.4
nf9 167.5 8.2 319 38 157.1 23.8 5 4.9024 41 12.1
nf10 191.6 21.5 327 39 169.7 44.1 5 4.9514 82.3 8.5
nf11 358.1 −9.7 310 35 347.5 −34 4 3.928 41.7 14.4
nf12 2.8 6.3 302 41 358.3 −29.1 4 3.4862 5.8 41.6
nf13 46.8 −13.7 285 23 52.5 −33 5 3.6377 2.9 54.4
nf15 170.8 18.9 39 5 169.3 17.7 4 3.9668 90.4 9.7
Mean 189 9.5 – – 179.3 30.5 8 7.2111 8.9 19.7
06 S (N) 175 23.6 290 15 171 37 8 7.8542 48 8
09 S 166 15 210 30 177 34 8 7.6667 21 13
09 N 5.2 −13.5 210 30 195 24 1 – – –
09 N + S 168 14.8 210 30 179 33 9 8.6190 21 12
13 N 339.9 −32.4 150 20 171 27 1 – – –
22 S 198.3 8.3 280 27 200 35 2 1.9873 79 11.2
22 N 34.1 −9.7 280 27 219 34 2 1.9655 29 18.5
22 N + S 205.7 9.2 280 27 209 35 4 3.9167 36 16.0
24 S 170.7 29.2 225 30 189 51 2 1.9412 17 24.4
67 N (S) 65.8 −4.2 24 95 194 41 23 22.6857 70 4.0
70 N 4.9 −26.5 216 46 216 40 2 1.9091 11 30.6
71 S 156.6 28.2 170 25 171 31 1 – – –
71 N 326.9 −28.1 170 25 162 35 4 3.9667 90 10.0
71 N + S 330.8 −28.8 170 25 166 34 5 4.9437 71 9.0
75 S 159 26.3 303 26 144 39 3 2.9259 27 15.8
75 N 19.3 1.2 303 26 198 24 5 4.9111 45 12
75 N + S 185.9 9.3 303 26 181 32 8 7.1250 8 20
76 S (N) 171.7 14.2 280 22 165 28 6 5.7826 23 14
80 N 334.1 −34.8 53 11 153 24 1 – – –
82 S 146.7 26.8 45 8 146 19 1 – – –
85 S 178.3 −5.4 250 45 183 37 10 9.9692 292 3
Mean 181 16.9 – – 179.7 33.7 29 27.1047 14.8 7.2

Figure 7. (a) Site mean directions in stratigraphic coordinates isolated through thermal demagnetization (all rotated to reverse polarity). (b) Comparison of
means between the formation mean from this study (solid circle) and the formation mean from Irving & Strong (1984) (solid square).

dried in a 45 mT field shows an anisotropy of 84 per cent with an the bootstrap technique of Constable & Tauxe (1990) to increase
RMS error of 1.7 and an a factor of 1.99. This a factor value is the sample number. Results from inclination corrections performed
consistent with previous observations for magnetite (e.g. Vaughn using a range of a factors (Bilardello 2008) indicate that the most
et al. 2005). In addition, ARM directions are closer to the applied symmetric and therefore geologically realistic distribution of site
field directions. The choice of a particle anisotropy of two was also mean VGPs lies between a factors of 1.4 and 2.2.
confirmed by evaluating the symmetry of corrected site mean VGPs Corrections performed using this range of a also agree, within
using the Bingham distribution (Bingham 1964) to analyse ellip- error, to an inclination correction performed using the EI tech-
tically distributed data (Beck 1999). VGPs were re-sampled using nique of Tauxe & Kent (2004) discussed earlier. This technique


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 283

Figure 8. Fold test performed on data from this study and Irving & Strong (1984) combined.

converts symmetric VGPs to directions, which is essentially similar 5 DISCUSSION


to evaluating the symmetry of corrected VGPs to determine particle
anisotropy. 5.1 Magnetic Mineralogy
Results of the thermal demagnetization of three orthogonal IRMs
(Lowrie 1990) suggest that haematite, magnetite and maghemite
4.5 Inclination correction
are present in the samples. The sharp drop observed at ∼680 ◦ C is
The uncorrected paleopole of the Deer Lake Group lies at 22.2◦ N, characteristic of haematite, however, this behaviour is not observed
122.3◦ E, A95 = 7.6◦ . When an inclination correction is applied using in the majority of our samples. All samples, however, show a loss
an a factor of 2 and the measured AAR fabric, the mean direction of intensity, even if slight, at ∼580 ◦ C, indicative of magnetite.
of the Deer Lake Group is corrected to D = 178.8◦ , I = 50.9◦ , The presence of maghemite is indicated by a slight loss of intensity
α 95 = 6.3◦ , N = 29, indicating an inclination shallowing of 17.2◦ between 350 and 450 ◦ C (Dunlop & Özdemir 1997; McElhinny
(remanent inclination measured, I = 33.7◦ ). When this direction is & McFadden 1999) and also so by the reapplication of an IRM at
converted to a paleopole for the Deer Lake Group, its position plots the end of the heating cycle, which shows an intensity much lower
at 8.4◦ N, 122.7◦ E, A95 = 7.2◦ , 13.8◦ S of the uncorrected paleopole. than before heating, assuming that the maghemite has inverted to
An inclination correction was also performed using the EI tech- haematite with heating. When the bulk susceptibility is plotted ver-
nique of Tauxe & Kent (2004), which is based on the TK03.GAD sus temperature (Fig. 10), a sudden increase is observed for cer-
model of the geomagnetic field using the software find EI (pmag- tain samples starting around 450 ◦ C, indicating that mineralogical
1.9.2 software package). This technique requires at least 100 sites to changes are occurring upon heating. The increased susceptibility,
represent the distribution of directions drawn from plausible models coupled with the reduced IRM capacity suggests transformation to
of the geomagnetic field (Tauxe & Kent 2004). Because only 29 site a phase with high IRM/χ (e.g. haematite) to one with very low
mean directions were available, we used specimen directions ob- IRM/χ (e.g. magnetite).
tained in this study together with the site mean directions reported Modelling of IRM acquisition curves (Kruiver et al. 2001)
by Irving & Strong (1984), which is justified because in sediments shows coercivity peaks that correspond to the coercivities of mag-
each specimen is a unique estimate of the geomagnetic field, giving netite/maghemite (50–70 mT) and haematite (500 mT), but also
58 sample directions. The correction yields a new inclination of higher coercivity peaks (2000 mT), which we interpret to be goethite
41.1◦ (33◦ and 52◦ , low and high 95 per cent confidence interval (Peters & Dekkers 2003) (Fig. 3).
inclinations), 9.8◦ smaller than the anisotropy-based correction, but MPMS low-temperature measurements confirm the results of the
with a large confidence interval that includes the previous result. IRM modelling. The loss of intensity measured on heating between


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
284 D. Bilardello and K. P. Kodama

Figure 9. Magnetic fabric in stratigraphic coordinates measured with AAR. (a) original data; (b) bootstrap data; (c) Flinn diagram (Flinn 1962) showing a
triaxial to oblate fabric. AAR magnetic fabric in geographic coordinates, (d) original data; (e) bootstrap data; (f) Histograms of the anisotropy eigevalues. Plots
were generated using pmag-1.9.2 (Tauxe 2006).

Figure 10. Bulk susceptibility as a function of temperature. Measurements where only possible for five sites for which we had spare samples. A sudden
increase of susceptibily at about 450 ◦ C was observed in three samples out of the five and is interpreted as an inversion of maghemite to haematite during
heating.

20 and 50 K, although not diagnostic, could be attributed to goethite, around 250 K. For pigmentary nanophase haematite, the Morin
as supported by the IRM acquisition modelling. Pyrrhotite also has a transition is reduced and shifted towards temperatures as low as
low-temperature transition at 34 K, however we believe the presence 5 K, until it is completely suppressed in particles ≤1 μm in size
of sulphides to be unlikely in such highly oxidized rocks. The sharp (Bando et al. 1965).
drop at 119 K is consistent with the Verwey transition (Dunlop and The FORC diagram, acquired in a peak field of 1.2 T, does
Özdemir 1997) that occurs in magnetite at these temperatures. The not show the high-coercivity phases, but rather only one low-
small fraction of remanence lost on cooling through the transition is coercivity group below 100 mT. This is probably because the
common for oxidized magnetite (Dunlop and Özdemir 1997). There higher magnetization of magnetite completely covers the signal of
is no indication of the Morin transition, diagnostic of haematite, haematite.


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 285

5.2 Paleomagnetism
The characteristic remanent magnetization (ChRM) is typically re-
vealed around 300 ◦ C and is completely unblocked by 600–625 ◦ C.
Beyond these temperatures, no stable component of the remanence
persists and it is not possible to isolate a direction.
AF demagnetization up to 50–60 mT isolated the same compo-
nents of magnetization as the thermal demagnetization, indicating
that the ChRM-carriers must be magnetic phases with low coerciv-
ities. Typically, these coercivities would be thermally demagnetized
by 580 ◦ C: magnetite has a Curie temperatures of 580 ◦ C, whereas
maghemite shows a loss of intensity in the 350–450 ◦ C range that
accompanies its inversion to haematite (Dunlop & Özdemir 1997;
McElhinny & McFadden 1999). Demagnetization of 1.2 T IRMs,
however, clearly reveals that haematite is present in the rocks, but
its natural remanence is for the most part not stable as shown in
the vector endpoint diagrams (Fig. 6). We therefore interpret the
ChRM to be carried by magnetite and/or maghemite. The ChRM
is only fully unblocked at temperatures of 600–625 ◦ C. Thermally
stable maghemite could, in theory, carry remanence up to its Curie
point ∼645 ◦ C (Dunlop & Özdemir 1997), alternatively, some finer
grained haematite could also be contributing to the remanence.
The magnetization of the Deer Lake Group thus appears to be a
pre-folding magnetization that has two antipodal, normal (north and Figure 11. Uncorrected and corrected paleopoles from the Maritime
negative) and reverse (south and positive), directions. In these red Provinces of Canada. Large open (filled) circle, uncorrected (corrected)
beds most of the ChRM is not carried by haematite of either detrital Deer Lake (UC and C) Group (this study); small filled circle, Deer Lake
or chemical origin, but by magnetite/maghemite, with fine grained Group (DL) (Irving & Strong 1984); Open (filled) diamond, uncorrected
haematite only partly contributing to the natural remanence. (corrected) Shepody Fm. (Shep UC and C) (Bilardello 2008; Bilardello &
Kodama 2010); Open (filled) triangle, uncorrected (corrected) Maringouin
Fm. Paleopoles (Mar UC and C); Grey shaded line is the fitted APW
5.3 Magnetic fabrics path for Laurentia from Torsvik et al. (1996) rotated into North Ameri-
can coordinates.
The fabric of magnetite/maghemite measured by AAR is sub-
parallel to bedding and would indicate a primary depositional fab-
ric. The fabric has its minimum axes steeply plunging to the NW. Formations. Because the amount of inclination shallowing is depen-
The maximum axes are clustered around the horizontal along a dent on the magnetic field inclination at deposition, which is in turn
NNE–SSW lineation, whereas the intermediate axes show a clus- a function of latitude, it is not surprising that the Deer Lake Group,
tering that gently plunges to the ESE. The magnetic lineation is deposited at higher latitudes with steeper initial inclinations, has
sub-parallel to the regional scale fold axes, suggesting that the fab- suffered from more inclination shallowing. Our results confirm the
ric could have been affected by strain. However, the beds that we need to move the Lower Carboniferous portion of North American’s
have sampled do not follow the trend of those structures, but dip APWP southward (Fig. 11) and strongly suggest a revision of the
predominantly to the NE and would not therefore record the same red bed-derived Paleozoic and Mesozoic portion of the APWP.
strain event. The southward translation of this portion of the APWP required
Excluding the tectonic fabric hypothesis, a depositional fabric by the corrections is ∼12◦ for the Middle Mississippian and ∼6◦
explanation remains a plausible interpretation for the formation of for the Upper Mississippian–Lower Pennsylvanian. As discussed in
the pronounced magnetic lineation. The minimum axes of the fab- Bilardello (2008) and Bilardello & Kodama (2010), this magnitude
ric, however, are very steep with the confidence ellipse including the translation would place Newfoundland in an arid climate zone dur-
pole to bedding and likewise, the maximum and intermediate axes ing the Middle Mississippian, consistent with the evaporitic deposits
are clustered on the horizontal, which means that the imbrication present in the Nova Scotia and New Brunswick Carboniferous rocks
of the fabric is not statistically significant. This could be the result (Geldsetzer et al. 1980). Newfoundland would then move progres-
of burial compaction leading to a bedding parallel fabric. We inter- sively northwards towards more humid conditions in the Middle to
pret the fabric as a depositional fabric, with good clustering of the Upper Pennsylvanian, as indicated by the thick coal deposits that
principal axes in stratigraphic coordinates, that has been affected by characterize the Cumberland and Pictou groups (Hacquebard 1970;
post-depositional compaction. Hacquebard & Donaldson 1969).
One other important effect of the southward translation of North
America (see Bilardello & Kodama 2010) is to increase the overlap
5.4 Inclination shallowing
between North America and Gondwana in the early Pangea assem-
Results of the inclination correction on the Deer Lake paleopole blages (Van der Voo 1993). Our results therefore favour a Pangea B
position are consistent with other inclination corrections performed assemblage (Muttoni et al. 2003, 2009), in which North America is
so far on the Maringouin and Shepody formations of the Maritime positioned to the west of Gondwana, thus eliminating the overlap.
Provinces of Canada (Bilardello 2008; Bilardello & Kodama 2010). In Bilardello & Kodama (2010), however, we argue that to fully
The 17.2◦ of correction for the Deer Lake Group is larger in magni- evaluate the effects of the correction it is necessary to consider
tude than the ∼10◦ of correction for the Shepody and Maringouin the placement of Gondwana for that time period. This is because


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
286 D. Bilardello and K. P. Kodama

the Carboniferous Gondwana paleopoles are derived primarily from between 0.4 and 0.77. Our results from haematite-bearing rocks of
sedimentary rocks (Torsvik & Van der Voo 2002), which are likely the Maritime Provinces of Canada imply f factors that range from
to have been affected by inclination shallowing. If this were the 0.64 to 0.83. The result of Tan et al. (2003) from Chinese Cretaceous
case an inclination correction would move Gondwana further south, red beds was also included.
possibly eliminating the overlap. We have deliberately left out values obtained from redeposition
experiments because when the redeposited sediments are measured
in the zero field of the magnetometer the remanent inclination may
5.5 Comparison with other inclination correction be affected by grains physically reorienting in the high water content
techniques sediments and therefore the measurement would not represent a true
King (1955) suggested a relationship between the measured (re- DRM (Tauxe & Kent 1984; Carter-Stiglitz et al. 2006). Likewise,
manent) inclination and field inclination to explain inclination shal- results of Tan et al. (2007) from the Newark Basin red beds have not
lowing observed during redeposition experiments. This relationship been included because Kent & Tauxe (2005) already include results
follows the equation: from the same Newark Basin Formations in their study.
Even higher inclination errors have been observed for Neogene
tan(lm ) = f tan(l0 ), (2) red sediments from the Vallès-Penedès basin (Garcés et al. 1996).
where I m is the measured, or remanent inclination; I 0 is the field Site mean inclinations of less than 20◦ are reported for three out
inclination and f is the shallowing factor that has been observed to of five finely laminated red bed sites, whereas they report a gamut
lie between 0.4 and 0.55 from experimental data (King 1955; Tauxe of inclinations ranging between 32◦ and 56◦ from six more sites
& Kent 1984). collected from finely laminated, burrowed, edafized and massive
We have collected inclination shallowing data for both magnetite facies. The expected geomagnetic field inclination is 60◦ . Garcés
and haematite from various sources and compiled it with our data et al. (1996) never report statistics at the Formation level, however,
from the Maritime Provinces of Canada (Table 2). All the results for calculation of a Formation mean (this study) from the 11 reported
magnetite are from anisotropy-based inclination shallowing correc- site means yielded D = 0.6◦ , I = 34.7◦ and α 95 = 10.2◦ .
tions, enabling direct calculation of the f factors, which range from Magnetite-bearing rocks display a range of f values between
0.54 (Deer Lake Group) to 0.79 (Nacimiento Fm). 0.54 and 0.79 with a mean value of 0.65 and a standard deviation
For haematite, Tauxe (2005) and Kent & Tauxe (2005) have stud- of 0.08 (N = 9). Haematite-bearing rocks have values between 0.4
ied Oligo-Miocene and mid Miocene haematite-bearing red sedi- and 0.83 with a mean of 0.59 and a standard deviation of 0.12
mentary rocks from Asia and Pakistan, and Late Triassic red beds (N = 15). These values correspond to greatly varying amounts of
from around the North Atlantic Ocean, respectively and have applied inclination shallowing for both magnetite- and haematite-bearing
an inclination shallowing correction using a statistical geomagnetic rocks (Fig. 12). It is very tempting to take the mean values and
field model (TK03.GAD), which relies on the elongation distribu- their standard errors (0.026 for magnetite and 0.031 for haematite)
tion of paleomagnetic directions caused by secular variation (Tauxe to perform simplified inclination corrections of formations that are
& Kent 2004). Their technique yielded f factors from both studies suspected to have suffered from inclination shallowing. However,

Table 2. Measured inclinations (I m ), corrected inclinations (I c ) and derived f factors for inclination
corrected magnetite- and haematite-bearing sedimentary rocks.
Magnetite and haematite f factors
Formation/group Magnetic carrier Im Ic f Reference
Glenshaw Fm. Magnetite 23.4 33.7 0.65 Kodama (2009)
Perforada Fm. Magnetite 35.6 50.1 0.60 Vaugn et al. (2005)
Nanaimo Group Magnetite 50.9 60.5 0.70 Kim & Kodama (2004)
Valle Group Magnetite 19.1 26.6 0.69 Li et al. (2004)
Ladd Fm. Magnetite 46 58 0.65 Tan & Kodama (1998)
Point Loma Fm. Magnetite 39.5 56 0.56 Tan & Kodama (1998)
Pigeon Point Fm. Magnetite 41.6 51.5 0.71 Kodama & Davi (1995)
Deer Lake Group Magnetite 33.7 50.9 0.54 This study
Nacimiento Fm. Magnetite 51.3 57.7 0.79 Kodama (1997)
Kapusaliang Fm. Haematite 29 52.5 0.43 Tan et al. (2003)
Shepody Fm. Haematite 20.4 30.1 0.64 Bilardello & Kodama (2010)
Maringouin Fm. Haematite 24.9 29.1 0.83 Bilardello & Kodama (2010)
Dan River Haematite 5.9 10 0.59 Kent & Tauxe (2005)
Newark basin: Princeton Haematite 5.2 9 0.57 Kent & Tauxe (2005)
Newark basin: Nursery Haematite 8.8 21 0.40 Kent & Tauxe (2005)
Newark basin: Titusville Haematite 13 20 0.63 Kent & Tauxe (2005)
Newark basin: Rutgers Haematite 14.2 21 0.66 Kent & Tauxe (2005)
Newark basin: Somerset Haematite 15.7 24 0.63 Kent & Tauxe (2005)
Newark basin: Weston Haematite 17.5 33 0.49 Kent & Tauxe (2005)
Newark basin: Martinsville Haematite 18.2 34 0.49 Kent & Tauxe (2005)
St. Audrie’s Bay Haematite 33.4 44 0.68 Kent & Tauxe (2005)
Jameson Land Haematite 45.1 60 0.58 Kent & Tauxe (2005)
Subei area Haematite 43.7 63 0.49 Tauxe (2005)
Potwar Plateau Haematite 33.7 41 0.77 Tauxe (2005)


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 287

Figure 12. Plot of the remanent (measured) inclinations (I m ) as a function of the field (corrected) inclinations (I 0 ). (a) Magnetite-bearing sedimentary rocks,
(b) is for haematite-bearing sedimentary rocks. The solid lines are the lines of perfect correlation between I m and I 0 (f = 1), the dashed lines comprise the
smallest and largest f factors measured, thus defining the ranges of maximum and minimum inclination shallowing observed. The dotted lines have been plotted
using the mean f factors. (a) The solid circle is from this study; the open circles are from other magnetic anisotropy-based studies (see Table 2). (b) The solid
circles are from our data (Bilardello 2008; Bilardello & Kodama 2010); the open circle in from the anisotropy-based correction of the Kapusaliang Fm.; open
triangles and diamonds are from the TK30.GAD corrections of Kent and Tauxe (2005) and Tauxe (2005) , respectively (see Table 2).

the f factor, from the inclination correction formula, incorporates of 0.59 yields an inclination 49.6◦ . This simple test indicates that
measurements of particle anisotropy and the magnetic fabric, which for these red beds the smallest of the observed f factors restores
make this parameter non-unique. Although particle anisotropy the inclination to that of the ambient geomagnetic field whereas the
may not vary substantially for haematite because it is not depen- mean or largest f factors observed would underestimate the correc-
dent on crystal shape (Dunlop & Özdemir 1997; Bilardello 2008; tion by ∼10◦ and ∼20◦ , respectively. Nevertheless, this test shows
Bilardello & Kodama 2009), it is for magnetite (Dunlop & Özdemir what ranges of inclination corrections may be expected for red beds
1997). Likewise the magnetic fabric of the rocks is dependent on that are suspected to be shallow.
the degree of alignment of the particles, which is expected to vary
with magnetic mineralogy and process (Jackson 1991). The range
of measured f factors is therefore more significant than the mean 6 C O N C LU S I O N S
value, when evaluating inclination errors. The inclination shallowing correction performed on the magnetite-
The maximum inclination error that can be expected from the bearing Deer Lake Group rocks using the measured AAR fabric
range of compiled f factors and the tan–tan relation of King (1955) and the measured particle anisotropy of the ChRM- carrying grains
is of 17.4◦ for a magnetite-bearing rock with an original inclination indicates 17.2◦ of shallowing. This amount of inclination shallow-
of 55◦ , using the smallest observed f factor of 0.54 (Fig. 12), and ing corresponds to a 13.8◦ increase in colatitude for the corrected
of 25.3◦ for a haematite-bearing rock with an original inclination of paleopole, a larger value than what has been obtained from other
60◦ , using the smallest observed f factor of 0.4 (Fig. 12). corrected red bed palepoles from the Maritime Provinces of Canada,
Inclination shallowing results may also be evaluated in terms but consistent with the tan–tan variation of measured, shallowed in-
of the dipole equation, tan I = 2tan λ, where λ is the latitude. clination as a function of field inclination. Together with the other
When evaluating I m versus the latitude calculated from the corrected inclination shallowing corrections that have been applied to this
inclinations, λ , the dipole relationship becomes portion of the Paleozoic and Mesozoic APWP for North America,
these results strongly suggest a revision of sedimentary rock, and
tan(Im ) =  tan(λ ), (3)
in particular red bed- derived, APWPs. The corrected paleolatitude
where  is a proportionality constant, the ‘paleolatitude correction’ puts Newfoundland in an arid climate zone for the Lower Carbonif-
constant. Because the field inclination versus remanent inclination erous, in agreement with lithologic indicators. The corrected paleo-
equation and the dipole equation have the same tan–tan form and are geography increases the overlap between North and South America
closely related, it is possible to calculate a corrected latitudinal range in a Pangea A configuration, making a Pangea B reconstruction
using  = 2f for sedimentary rocks whose measured remanent necessary for this time period.
inclinations are suspected to be shallow. However, to fully evaluate the effects of the correction it is nec-
To test these relationships, we have applied simplified inclination essary to consider the positioning of Gondwana.
corrections to our calculated mean (D = 0.6◦ , I = 34.7◦ , α 95 = The distribution of the corrected paleomagnetic directions pro-
10.2◦ , N = 11) of the Vallès-Penedès basin red beds from the data vides a means of estimating the a factor used for the anisotropy
of Garcés et al. (1996) (expected field inclination = 60◦ ). correction, and results suggest that the corrected inclinations corre-
When we apply a correction using the observed haematite shal- spond to a geologically reasonable distribution of directions.
lowing factor end members of 0.4 and 0.83, the corrected inclina- A comparison of the proportionality constant between the tan-
tions are 60◦ and 39.8◦ . A correction applied using the mean f factor gent of the remanent inclinations and the tangent of the field


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
288 D. Bilardello and K. P. Kodama

inclinations, the flattening factor f , derived from our anisotropy- Constable, C.G. & Tauxe, L., 1990. The bootstrap for magnetic susceptibility
based inclination corrections and f factors resulting from other tensors, J. geophys. Res., 95(B6), 8383–8395.
studies show strong agreement. A compilation of f factors from in- Deamer,G.A. & Kodama, K.P. 1990. Compaction-induced inclination shal-
clination corrections performed on both magnetite- and haematite- lowing in synthetic and natural clay-rich sediments, J. geophys. Res.,
95(B4), 4511–4529.
bearing rock formations allowed to determine the range of f factors
Dunlop, D.J. & Özdemir, Ö., 1997. Rock Magnetism: Fundamentals and
that may be expected, allowing determination of the possible ranges
frontiers, Cambridge Studies in Magnetism, 573 pp., Cambridge Univer-
of inclination shallowing. sity Press, Cambridge.
The Neogene Vallès-Penedès basin red beds have been corrected Elston, D.P. & Purucker, M.E., 1979. Detrital magnetization in red beds of
to test the range of derived f factors. The smallest of the f fac- the Moenkopi Formation (Triassic) Gray Mountain, Arizona, J. geophys.
tors perfectly restores the inclination to that of the ambient field, Res., 84(B4), 1653–1665.
whereas the mean and largest f factors considerably undercorrect Flinn, D., 1962. On folding during three-dimensional progressive deforma-
the inclination. This simple test indicates what ranges of inclination tion, Q. J. Geol. Soc. Lond., 118, 385–428.
corrections may be expected for sedimentary rocks suspected to be Garcés, M., Parés, J.M. & Cabrera, L., 1996. Inclination error linked to sed-
shallow. imentary facies in Miocene detrital sequences from the Vallès-Penedès
Basin (NE Spain), in Palaeomagnetism and tectonics of the Mediter-
ranean Region, Vol. 105, pp. 91–99, eds. Morris, A. & Tarling, D.H.
Geological Society Special Publication.
Geldsetzer, H.H.J., Giles, P., Moore, R. & Palmer, W., 1980. Stratigraphy,
AC K N OW L E D G M E N T S Sedimentology and Mineralization of the Carboniferous Windsor Group,
Nova Scotia, GAC/MAC Field Trip Guidebook No. 22, Department of
The research was supported by National Science Foundation grant
Geology, Dalhousie University, Halifax.
EAR-540204 to K.P. Kodama. The authors thank Dennis V. Kent
Gilder, S., Chen, Y. & Sen, S., 2001. Oligo-Miocene magnetostratigraphy
for initial comments on the manuscript, Amy P. Chen for the FORC and rock magnetism of the Xishuigou section, Subei (Gansu Province,
diagram hysteresis measurements and Michael Winklhofer for his western China) and implications for shallow inclinations in central Asia,
expertise on FORC data and discussion. Lisa Tauxe and Michael J. J. geophys. Res., 106(B12), 12, 30 505–30 521.
Jackson substantially improved the manuscript with their thoughtful Hacquebard, P.A., 1970. Coal in Southeastern Canada, Geol. Survey of
comments and reviews. Canada, Economic Geology Report 1.
Hacquebard, P.A. & Donaldson, J.R., 1969. Carboniferous coal deposition
associated with floodplain and limnic environments in Nova Scotia, Geol.
REFERENCES Soc. Am. Special Paper, 114, 143–191.
Hounslow, M.W. & Maher, B.A., 1999. Laboratory procedures for quan-
Anson, G.L. & Kodama, K.P., 1987. Compaction-induced inclination shal- titative extraction and analysis of magnetic minerals from sediments,
lowing of the post-depositional remanent magnetization in a synthetic in Environmental Magnetsm: A Practial Guide, pp. 139–184, eds. J.
sediment, Geophys. J. R. astr. Soc., 88, 673–692. Walden, F. Oldfield & J.P. Smith, Quaternary Research Association,
Bando, Y., Kiyama, M., Yamamoto, N., Takada, T., Shinjo, T. & Takaki, H., Cambridge.
1965. The magnetic properties of α-Fe2 O3 fine particles, J. Phys. Soc. Hyde, R.S., 1979. Geology of the Carbonifeorous strata in portions of the
Japan, 20, 2086. Deer Lake Basin, western Newfoundland, Mineral Development Divi-
Beck, M.E., 1999. On the shape of paleomagnetic data sets, J. geophys. Res. sion, Department of Mines and Energy Government of Newfoundland
104(B11), 25 427–25 441. and Labrador, 79(6), 43.
Belt, E.S., 1969. Newfoudland Carbonifeous stratigraphy and its relation Hyde, R.S., 1984. Geology and mineralization of the Carboniferous Deer
to the Maritimes and Ireland, Am. Assoc. Petrol. Geol. Mem., 12, 734– Lake basin, western Newfoundland in Mineral Deposits of Newfound-
753. land: a 1984 Perspective, Report 84-3, pp. 19–26, Mineral Development
Bilardello, D., 2008. A new technique for measuring the magnetic fabric of Division, Government of Newfoundland and Labrador, St. John’s, NL.
hematite-bearing sedimentary rocks, hf-AIR: inclination correction case Hyde, R.S. & Ware, M.J., 1981. Notes on the geology of the Deer Lake
studies of Carboniferous red beds from the Maritime Provinces of Canada, (12H/3) and Rainy Lake (12A/14) Map Areas, Newfoundland Department
PhD thesis, Lehigh University, Bethlehem, PA. Mines Energy, Mineral Development Division, 7 pp.
Bilardello, D. & Kodama, K.P., 2009a. Measuring remanence anisotropy of Irving, E. & Strong, D.F., 1984. Paleomagnetism of the Early Carbonif-
hematite in red beds: anisotropy of high-field isothermal remanence mag- erous Deer Lake Group, western Newfoundland: no evidence for mid-
netization (hf-AIR), Geophys. J. Int., 178, 1260–1272, doi:10.111/j.1365- Carboniferous displacement of “Acadia”, Earth planet. Sci. Lett., 69,
246X.2009.04231.x. 379–390.
Bilardello, D. & Kodama, K.P., 2010. Paleomagnetism and magnetic Jackson, M., 1991. Anisotropy of magnetic remanence: a brief review of
anisotropy of Carboniferous red beds from the Maritime Provinces of mineralogical sources, physical origins, and geological applications, and
Canada: evidence for shallow paleomagnetic inclinations and implica- comparison with susceptibility anisotropy, Pure. appl. Geophys., 136(1),
tions for North American apparent polar wander, Geophys. J. Int., 180, 1–28.
1013–1029, doi:10.1111/j.1365-246X.2009.04457.x. Jackson, M., Banerjee, S.K., Marvin, J.A., Lu, R. & Gruber, W., 1991. Detri-
Bingham, C., 1964. Distributions on the sphere and on the projective plane, tal remanence inclination errors and anhysteretic remanence anisotropy:
PhD thesis, Yale University. quantitative model and experimental results, Geophys. J. Int., 104,
Blow, R.A., & Hamilton, N., 1978. Effect of compaction on the acquisition 95–103.
of a detrital remanent magnetization in fine-grained sediments, Geophys. Kent, D.V. & Tauxe, L., 2005. Corrected late Triassic latitudes for
J. R. Astr. Soc., 52, 13–23. continents adjacent to the North Atlantic, Science, 307, 240–244,
Carter-Stiglitz, B., Valet, J.P. & LeGoff, M., 2006. Constraints on the acqui- doi:10.1126/science.1105826.
sition of remanent magnetization in fine-grained sediments imposed by Kim, B.Y. & Kodama, P.K., 2004. A compaction correction for the pale-
redeposition experiments, Earth planet Sci. Lett., 245, 427–437. omagnetism of the Nanaimo Group sedimentary rocks: implications for
Collinson, D.W., 1965. The remanent magnetization and magnetic properties the Baja British Columbia hypothesis, J. geophys. Res., 109(B02102),
of red sediments, Geophys. J. R. astr. Soc., 10, 105–126. doi:10.1029/2003JB002696.
Collinson, D.W., 1974. The role of pigment and specularite in the remanent King, R.F., 1955. The remanent magnetism of artificially deposited sedi-
magnetism of red sandstone, Geophys. J. R. Astr. Soc., 38, 253–264. ments, Mon. Notic. Roy. Astr. Soc. geophys. Suppl., 7, 115–134.


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS
Rock magnetism and inclination shallowing 289

Kodama, K.P., 1997. A successful rock magnetic technique for correcting Sun, W.W. & Kodama, K.P., 1992. Magnetic anisotropy, scanning elec-
paleomagnetic inclination shallowing: Case study of the Nacimiento For- tron microscopy, and X ray pole figure goniometry study of inclina-
mation of New Mexico, J. geophys. Res., 102(B3), 5193–5206. tion shallowing in a compacting clay-rich sediment, J. geophys. Res., 97,
Kodama, K.P., 2009. Simplification of the anisotropy-based inclination cor- 9599–9615.
rection technique for magnetite- and hematite bearing rocks: a case Tan, X. & Kodama, K.P., 1998. Compaction-corrected inclinations from
study for the Carboniferous Gleshaw and Mauch Chunk Formations, southern California Cretaceous marine sedimentary rocks indicate no
North America, Geophys. J. Int., 176, 467–477, doi:10.1111/j.1365- paleolatitudinal offset for the Peninsular Ranges terrane, J. geophys. Res.,
246X.2008.04013.x. 103(B11), 27 169—27 192.
Kodama, K.P. & Davi, J.M., 1995. A compaction correction for the pa- Tan, X. & Kodama, K.P., 2002. Magnetic anisotropy and paleomagnetic
leomagnetism of the Cretaceous Pigeon Point Formation of California, inclination shallowing in red beds: evidence from the Mississippian
Tectonics, 14(5), 1153–1164. Mauch Chunk Formation, Pennsylvania, J. geophys. Res., 107(B11), 2311,
Kodama, K.P. & Dekkers, M.J., 2004. Magnetic anisotropy as an aid to doi:10.1029/2001JB001636.
identifying CRM and DRM in red sedimentary rocks, Stud. geophys. Tan, X. & Kodama, K.P., 2003. An analytical solution for correcting paleo-
Geod., 48, 747–766. magnetic inclination error, Geophys. J. Int., 152, 228–236.
Kruiver, P.P., Dekkers M.J. & Heslop, D., 2001. Quantification of magnetic Tan, X., Kodama, K.P., Chen, H., Fang, D., Sun, D. & Li, Y., 2003. Paleomag-
coercivity components by the analysis of acquisition curves of isothermal netism and magnetic anisotropy of Cretaceous red beds from ther Tarim
remanent magnetization, Earth planet. Sci. Lett., 189, 269–276. basin, northwest China: Evidence for a rock magnetic cause of anoma-
Larson, E.E., Walker, T.R., Patterson, P.E., Hoblitt, R.P. & Rosenbaum, J.C., lously shallow paleomagnetic inclinations from central Asia, J. geophys.
1982. Paleomagnetism of the Moenkopy Formation, Colorado Plateau: Res., 108(B2), 2107, doi:10.1029/2001JB001608.
basis for long term model of acquisition CRM in red beds, J. geophys. Tan, X., Kodama, K.P., Gilder, S. & Courtillot, V., 2007. Rock magnetic
Res., 87, 1081–1106. evidence for inclination shallowing in the Passaic Formation red beds
Li, Y., Kodama, K.P. & Smith, D.P., 2004. New paleomagnetic, rock magnetic from the Newark basin and a systematic bias of the Late Triassic apparent
and petrographic results from the Valle Group, Baja California, Mexico: polar wander path for North America, Earth planet. Sci. Lett., 254(3–4),
exploring the causes of anomalously shallow paleomagnetic inclinations, 345–357.
J. geophys. Res., 109(B11101), doi:10.1029/2004JB003127. Tauxe, L., 2005. Inclination flattening and the geogentric axial dipole hy-
Liebes, E. & Shive, P. N., 1982. Magnetic acquisition in two Meso- pothesis, Earth planet. Sci. Lett., 233, 247–261.
zoic red sandstones, Phys. Earth planet. Sci. Inter., 30(4), 396–404, Tauxe, L., 2006. pmag- 1.9.2, http://magician.ucsd.edu/Software/PMAG/.
doi:10.1016/0031-9201(82)90049-8. Tauxe, L., Constable, C., Stokking, L. & Badgley, C., 1990. Use of
Lowrie, W., 1990. Identification of ferromagnetic minerals in a rock by anisotropy to determine the origin of characteristic remanence in the
coercivity and unblocking temperature properties, Geophys. Res. Lett., Siwalik red beds of northern Pakistan, J. geophys. Res., 95(B4), 4391–
17, 159–162. 4404.
McCabe, C., Jackson, M. & Elwood, B.B., 1985. Magnetic anisotropy in the Tauxe, L. & Kent, D.V., 1984. Properties of a detrital remanence carried by
Trenton limestone: Results of a new technique, anisotropy of anhysteretic hematite from study of modern river deposits and laboratory redeposition
susceptibility, Geophys. Res. Lett., 12(6), 333–336. experiments, Geophys. J. R. Astr. Soc., 77, 543–561.
McElhinny, M.W. & Lock, J., 1993. Global paleomagnetic database supple- Tauxe, L. & Kent, D.V., 2004. A new statistical model for the geomag-
ment number one, update to 1992, Surv. Geophys., 14, 303–329. netic field and the detection of shallow bias in paleomagnetic inclina-
McElhinny, M.W. & Lock, J., 1996. IAGA paleomagnetic databases with tions: was the ancient field dipolar? in Timescales of the Internal Geo-
Access, Surv. Geophys., 17, 575–591. magnetic Field, Vol. 145, pp. 101–116, J.E.T. Channell, D. V. Kent, W.
McElhinny, M.W. & McFadden, P.L., 1999. Paleomagnetism: Continents Lowrie and J. Meert, eds., Geophysics Monograph, American Geophysics
and Oceans, International Geophysics Series, 73, 385 pp., Academic Union.
Press, San Diego, USA. Tauxe, L. & Watson, G., 1994. The fold test: an eigen analysis approach,
McFadden, P.L. & Jones, D.L., 1981. The fold test in paleomagnetism, Earth planet. Sci. Lett., 122, 331–341.
Geophys. J. R. Astr. Soc., 67, 53–58. Torsvik, T. & Van Der Voo, R., 2002. Refining Gondwana and Pangea
Mitra R. & Tauxe, L., 2009. Full vector model for magnetization in sedi- palaeogeography: estimates of Phanerozoic non-dipole (octupole) fields,
ments, Earth planet. Sci. Lett., 286, 535–545. Geophys. J. Int., 151, 771–794.
Muttoni, G., Kent, D.V., Garzanti E., Brack P., Abrahamsem N. & Gaetani, Van Der Voo, R., 1990. Phanerozoic paleomagnetic poles from Europe and
M., 2003. Early Pangea ‘B’ to Late Permian Pangea ‘A’, Earth planet. North America and comparisons with continental reconstructions, Rev.
Sci. Lett., 215, 379–394. Geophys., 28, 167–206.
Muttoni, G. et al., 2009. Opening of the Neo-Tethys Ocean and the Pangea Van Der Voo, R., 1993. Paleomagnetism of the Atlantic, Tethys and Iapetus
B to Pangea A transformation during the Permian, GeoArabia, 14(4), oceans, 411 pp., Cambridge University Press, Cambridge.
17–48. Vaughn, J., Kodama, K.P. & Smith, D.P., 2005. Correction of inclination shal-
Peters, C. & Dekkers, M.J., 2003. Selected room temperature magnetic lowing and its tectonic implications: the Cretaceous Perforada Formation,
parameters as a function of mineralogy, concentration ad grain size, Phys. Baja California, Earth planet. Sci. Lett., 232, 71–82.
Chem. Earth, 28, 659–667. Walker, T.R., Larson, E.E. & Hoblitt, R.P., 1981. Nature and origin of
Rees, A.I., 1966. The effect of depositional slopes on the anisotropy of mag- hematite in the Moenkopy Formation (Triassic), Colorado Plateau: a con-
netic susceptibility of laboratory deposited sands, J. Geol., 74, 856–867. tribution to the origin of magnetism in red beds, J. geophys. Res., 86,
Roy, J.R. & Park, J.K., 1972. Red beds: DRM or CRM? Earth planet Sci. 317–333.
Lett., 17, 211–216. Winklhofer, M., & Zimanyi G.T., 2006. Extracting the intrinsic switching
Scherbakov, V. & Scherbakova, V., 1983. On the theory of depositional re- field distribution in perpendicular media: a comparative analysis, J. appl.
manent magnetization in sedimentary rocks, Geophys. Surv., 5, 369–380. Phys., 99, 08E710.
Steiner, M.B., 1983. Detrital remanent magnetization in hematite, J. geophys. Zijderveld, J.D.A., 1967. A.C. demagnetization of rocks, in Methods in
Res., 88, 6523–6539. Palaeomagnetism, pp. 256–286, eds. D.W. Collinson, K.M. Creer and
Stephenson, A., Sadikun, S. & Potter, D.K., 1986. A theoretical and ex- S.K. Runcorn, Elsevier, New York.
perimental comparison of the anisotropies of magnetic susceptibility and Zijderveld, J.D.A., 1975. Paleomagnetism of the Esterel rocks, PhD thesis,
remanence in rocks and minerals, Geophys. J. R. Astr. Soc., 84, 185–200. 199 pp., State University Utrecht, Holland.


C 2010 The Authors, GJI, 181, 275–289
Journal compilation 
C 2010 RAS

Potrebbero piacerti anche